You are on page 1of 24

MEASURES INDUCED BY UNITS

GIOVANNI PANTI AND DAVIDE RAVOTTI


arXiv:1203.2513v3 [math.LO] 10 Jan 2013

Abstract. The half-open real unit interval (0, 1] is closed under the ordinary
multiplication and its residuum. The corresponding infinite-valued proposi-
tional logic has as its equivalent algebraic semantics the equational class of can-
cellative hoops. Fixing a strong unit in a cancellative hoop —equivalently, in
the enveloping lattice-ordered abelian group— amounts to fixing a gauge scale
for falsity. In this paper we show that any strong unit in a finitely presented
cancellative hoop H induces naturally (i.e., in a representation-independent
way) an automorphism-invariant positive normalized linear functional on H.
Since H is representable as a uniformly dense set of continuous functions on its
maximal spectrum, such functionals —in this context usually called states—
amount to automorphism-invariant finite Borel measures on the spectrum. Dif-
ferent choices for the unit may be algebraically unrelated (e.g., they may lie
in different orbits under the automorphism group of H), but our second main
result shows that the corresponding measures are always absolutely continuous
w.r.t. each other, and provides an explicit expression for the reciprocal density.

1. Introduction
· 0) satisfying
A cancellative hoop is an algebra (H, +, −,
commutative monoid identities for +, 0,
x−· x = 0,
· y) −
(x − · z=x− · (y + z),
· · y),
x + (y − x) = y + (x −
· y = x.
(x + y) −
The above identities have manifold appearances, and can be seen from various
points of view.
First of all, they characterize the equational theory of truncated addition. In-
deed, the class CH of cancellative hoops coincides with the equational class gen-
· 0), where x −
erated by (R≥0 , +, −, · y = max{0, x − y}. Therefore, they suffice to
capture all identities valid in the positive reals. Also, cancellative hoops are pre-
cisely the positive cones of lattice-ordered abelian groups (`-groups for short), the
latter being the abelian groups (G, +, −, 0) endowed with a lattice order (G, ∧, ∨)
on which the group translations act as order automorphisms. Actually, more is
true: the assignment G 7→ G≥0 = {g ∈ G : g ≥ 0} is functorial, and provides a cat-
egorical equivalence between the class of all `-groups (which is generated by R, and
is equational in the language +, −, 0, ∧, ∨), and CH [6, Theorem 1.17]. Note that
· x) and
the lattice operations are definable in the hoop language by x ∨ y = x + (y −

Key words and phrases. cancellative hoop, lattice-ordered abelian group, MV-algebra, strong
unit, uniform distribution, Cesàro mean.
2010 Math. Subj. Class.: 03G25; 06F20; 11K06.
1
2 G. PANTI, D. RAVOTTI

x∧y = x − · (x − · y). The prototypical examples of `-groups —and hence, taking


positive cones, of cancellative hoops— are the sublattice-subgroups of the `-group
C(X) of all real-valued continuous functions on a compact Hausdorff space X, with
componentwise operations.
Reversing the order and changing notation provides a second point of view, of
a logical nature. Let exp(r) = cr , for some arbitrarily fixed real number 0 <
c < 1. Then exp is an order-reversing isomorphism between (R≥0 , +, −, · 0) and
the residuated lattice ((0, 1], ·, →, 1); here (0, 1] is the half-open real unit interval
endowed with the usual order and multiplication ·, and → is the residuum of ·,
characterized by the adjunction z ≤ x → y iff z · x ≤ y. An easy checking gives
x → y = min{1, y/x}, and the cancellative hoop axioms translate to the familiar
logical identities

commutative monoid identities for ·, 1,


x → x = 1,
z → (y → x) = (z · y) → x,
x · (x → y) = y · (y → x),
y → (x · y) = x.

The many-valued logical system defined by the above identities has the interval
(0, 1] as its set of truth-values, and is called the logic of cancellative hoops [11],
[26], [21, §6.4]; it is what remains of Hajek’s product logic [15] once the falsum is
removed.
A third facet of the matter arises from the observation that the addition opera-
tion + and its residuum − · capture the process of production and consumption of
resources in a Petri net under the firing of a transition. In the classical setting the
resources at each place of a Petri net are marked by an element of Z≥0 , and this
setting is readily generalizable to arbitrary cancellative hoops which are called, in
this setting, Petri algebras. See [1] and references therein.
An element of the cancellative hoop H whose multiples eventually dominate any
element of H is traditionally called a strong unit or, sometimes, an order unit.
Since we are not dealing with the related notion of weak unit, we will simply say
unit. So, explicitly, a unit is an element u ∈ H such that for every h ∈ H there
exists a positive integer k with h ≤ ku. An element u of a hoop of the form C(X)≥0
is a unit iff u(x) 6= 0 for every x ∈ X (remember that X is compact Hausdorff by
definition). Units arise in all contexts cited above: they are archimedean elements
in the `-group enveloping H, gauge scales for falsity in falsum-free product logic,
and universal bounds in bounded Petri nets. Upon defining x ⊕ y = u ∧ (x + y),
¬x = u − x, x y = ¬(¬x ⊕ ¬y), the interval Γ(H, u) := {h ∈ H : h ≤ u} is an MV-
algebra, namely the equivalent algebraic semantics of Lukasiewicz infinite-valued
logic. In the following we assume some familiarity with `-groups, cancellative hoops
and MV-algebras, referring to [5], [6], [8], [11], [24] for all unproved claims.
Let u be a unit in the cancellative hoop H. A state on the pair (H, u) is a
monoid homomorphism m : H → R≥0 such that m(u) = 1; states on (H, u) corre-
spond in a 1–1 canonical way to states on the unital `-group enveloping H and to
states on the MV-algebra Γ(H, u) [22, Theorem 2.4]. We are naturally interested
in automorphism-invariant states, namely those states m such that m ◦ σ = m for
every automorphism σ of H that fixes u. This is a quite natural requirement, since
MEASURES INDUCED BY UNITS 3

its lacking makes impossible to assign an “average truth value” to propositions in


infinite-valued Lukasiewicz logic in a representation-independent way. The first re-
sult in this context is [22, Theorem 3.4], saying that integration w.r.t. the Lebesgue
measure is an automorphism-invariant state on the free MV-algebra over finitely
many generators Freen (MV); this has been generalized in [23, Theorem 4.1] to all
finitely presented MV-algebras. As a matter of fact, integration w.r.t. the Lebesgue
measure is the only (modulo certain natural restrictions) automorphism-invariant
state on Freen (MV); this has been proved in [27] using ergodic theory, and in [20]
using algebraic means.
The set Mu of all states on (H, u) is compact convex in RH ; let Eu ⊆ Mu be
the set of extremal states. Then m ∈ Eu iff m is a hoop homomorphism [14,
Theorem 12.18]. As in [14, p. 70], we say that e ∈ Eu is discrete if e[H] is a
discrete subhoop of R≥0 : this happens iff e[H] = b−1 · Z≥0 for a uniquely defined
integer b ≥ 1, which we call the denominator of e. Let H be any finitely presented
cancellative hoop, and let u ∈ H be any unit. Our first main result, Theorem 4.2,
says that for every enumeration of the discrete extremal states of (H, u) according
to nondecreasing denominators, the corresponding Cesàro mean converges, does
not depend on the enumeration, and determines an automorphism-invariant state
mu on (H, u). The definition of mu is thus intrinsic, because it is not based on any
given representation of H. Of course the proof of the above properties is heavily
based on the representation theory of `-groups (as well as on tools new in this circle
of ideas, namely uniform distribution theory and the Ehrhart theory).
Let now a representation of H as a separating subhoop of some C(X)≥0 be
given. Contrary to standard usage, we do not assume that the elements of H are
represented by piecewise-linear functions on X, nor that u is represented by the
constant function 1l (note however that the representing space X is determined
by H up to homeomorphism; see §2). By Lemma 2.1(iv) the state mu defined by
Theorem 4.2 corresponds to a unique Borel finite measure µu on X. Let now v be
any other unit of H, and let mv , µv be the corresponding state and measure. In
contrast with the vector lattice case [4, Lemma 4.2], the automorphism group of
H does not act transitively on the set of units, so u and v may be algebraically
unrelated. Nevertheless, our second main result, Theorem 6.1, says that µu and µv
are absolutely continuous w.r.t. each other, and computes explicitly the reciprocal
density in terms of u, v, and the dimension of X —again, a number intrinsically
determined by H.
We thank the referee for his attentive reading and detailed comments on a pre-
vious version of this paper.

2. Representations and relative volume


We fix notation by recalling a few basic definitions and facts; since the categorical
equivalence between `-groups and cancellative hoops preserves all the usual notions
of homomorphism, subalgebra, dual space, . . ., we shall use freely well-known facts
from the theory of `-groups.
The maximal spectrum of the cancellative hoop H is the space MaxSpec H whose
elements are the maximal ideals of H (i.e., the kernels of the hoop homomorphisms
from H to R≥0 ), endowed with the hull-kernel topology, namely the topology gener-
ated by all sets of the form {m ∈ MaxSpec H : h ∈ / m}, for h varying in H. All hoops
we consider are representable in the following restricted sense: H is representable
4 G. PANTI, D. RAVOTTI

if it is nonzero and there exists an injective hoop homomorphism ρ : H → C(X)≥0 ,


for some compact Hausdorff space X, such that for every x 6= y ∈ X there exists
h ∈ H with (ρh)(x) = 0 and (ρh)(y) 6= 0.
Lemma 2.1. Let H be a cancellative hoop. Then:
T
(i) H is representable iff it contains a unit and MaxSpec H = 0.
(ii) Let ρ : H → C(X)≥0 be a representation and u ∈ H a unit. Then the three
spaces Eu , MaxSpec H, X are canonically homeomorphic via the mappings
X 3 x 7→ mx = {h ∈ H : (ρh)(x) = 0} ∈ MaxSpec H;
MaxSpec H 3 m 7→ the unique hoop homomorphism e : H → R≥0

whose kernel is m and such that e(u) = 1 ∈ Eu ;

X 3 x 7→ h 7→ (ρh)(x)/(ρu)(x) ∈ Eu .
(iii) Given two representations ρi : H → C(Xi )≥0 , for i = 1, 2, let R : X1 →
X2 be the homeomorphism obtained by composing the maps in (ii) (this
amounts to saying that Rx is the unique point in X2 such that, for every
h ∈ H, (ρ1 h)(x) = 0 iff (ρ2 h)(Rx) = 0). Then there exists a unit f ∈
C(X1 )≥0 such that, for every h ∈ H,

ρ1 h = f · (ρ2 h) ◦ R .
Moreover, R is the only homeomorphism : X1 → X2 satisfying this prop-
erty.
(iv) Let ρ, u be as in (ii). Then ρ induces a canonical bijection between the set
of states on (H, u) and the set of regular Borel measures on X (since X is
compact, such measures are necessarily finite). If H is countable, all Borel
measures on X are regular.
Proof. (i) and (ii) are well known [33].
(iii) Let x ∈ X1 ; then {(ρ1 h)(x) : h ∈ H} and {(ρ2 h)(Rx) : h ∈ H} are
isomorphic nonzero subhoops of R≥0 under the natural isomorphism (since they
both are isomorphic to H/mx ). By [17, Proposition II.2.2], any isomorphism of
subhoops of R≥0 is given by multiplication by a positive real  number. Therefore
there exists f (x) > 0 such that (ρ1 h)(x) = f (x)· (ρ2 h)(Rx) , for every h ∈ H. The
function x 7→ f (x) is continuous because it is the quotient f = (ρ1 u)/ (ρ2 u) ◦ R
of two continuous never 0 functions (u being any unit of H). Since f is never 0
and X1 is compact, f is a unit of C(X1 )≥0 ; in general f does not belong to ρ1 H.
Let T : X1 → X2 be a homeomorphism suchthat, for some unit g of C(X1 )≥0 and
every h ∈ H, we have ρ1 h = g · (ρ2 h) ◦ T . Let x ∈ X1 ; since (ρ1 h)(x) = 0 iff
(ρ2 h)(T x) = 0 for every h ∈ H, we have T x = Rx.
(iv) Every state m on (H, u) obviously corresponds to the state m ◦ ρ−1 on
(ρH, ρu). As proved in [27, Proposition 1.1] using the lattice version of the Stone-
Weierstrass theorem, m ◦ ρ−1 can be uniquely extended to a state on all of C(X).
By the Riesz representation theorem there exists a 1–1 correspondence between
states on C(X) and Borel regular measures on X; denoting with µ the measure
corresponding to m ◦ ρ−1 we thus have
Z
−1
m(h) = (m ◦ ρ )(ρh) = ρh dµ,
X
MEASURES INDUCED BY UNITS 5

for every h ∈ H. If H is countable then X is second countable, whence metrizable


and every open set is σ-compact; hence every Borel measure on X is regular [30,
Theorem 2,18]. See [18], [27], [9], [13], [24, Chapter 10] for details and further
developments. 
A McNaughton function over the n-dimensional cube [0, 1]n is a continuous func-
tion f : [0, 1]n → R≥0 for which the following holds:
there exist finitely many affine polynomials f1 , . . . , fk , each fi of the form
fi = a1i x1 + a2i x2 + · · · + ani xn + an+1
i , with a1i , . . . , an+1
i integers, such that,
n
for each w ∈ [0, 1] , there exists i ∈ {1, . . . , k} with f (w) = fi (w).
The free cancellative hoop over n + 1 generators Freen+1 (CH) is then the sub-
hoop M ([0, 1]n )≥0 of C([0, 1]n )≥0 whose elements are all the McNaughton func-
tions, the free generators being x1 , . . . , xn , 1l − (x1 ∨ · · · ∨ xn ); here xi is the
i-th projection [26, Theorem 1]. Also, the free MV-algebra over n generators is
Freen (MV) = Γ(Freen+1 (CH), 1l), the free generators being x1 , . . . , xn . Extending
the above notation, we will write M (W )≥0 for the hoop of restrictions of Mc-
Naughton functions to the closed subset W of [0, 1]n .
We write points of [0, 1]n as n-tuples w = (α1 , . . . , αn ), and points of Rn+1 as
column vectors w = (α1 · · · αn+1 )tr . We also use boldface to denote functions on
subsets of Rn+1 ; in particular x1 , . . . , xn+1 are the coordinate projections. We often
embed [0, 1]n in the hyperplane {xn+1 = 1} of Rn+1 in the obvious way. The point
w is rational if it belongs to Qn+1 , and is integer if it belongs to Zn+1 . In the
latter case, w is primitive if the gcd of its coordinates is 1. The denominator of
w ∈ Qn+1 is the least integer b ≥ 1 such that bw ∈ Zn+1 .
The cone over W ⊆ [0, 1]n ⊂ Rn+1 is Cone(W ) = {αw : α ∈ R>0 and w ∈ W }.
Any f ∈ M (W )≥0 gives rise to its homogeneous
 correspondent f : Cone(W ) → R≥0
via f (w) = xn+1 (w) · f w/xn+1 (w) , which is a positively homogeneous piecewise-
linear map, all of whose linear pieces have integer coefficients. In syntactical terms,
f is obtained by writing f as a term (either in the `-group language +, −, 0, ∨, ∧
or in the hoop language) built up from the projections x1 , . . . , xn and the constant
function 1l, and replacing each xi with xi , and 1l with xn+1 .
Definition 2.2. A McNaughton representation of the cancellative hoop H is a
representation ρ : H → C(W )≥0 such that W is a closed subset of some cube
[0, 1]n and the range of ρ is M (W )≥0 . We then write h : Cone(W ) → R≥0 for the
homogeneous correspondent of ρh. Given a unit u of H, we set W1 = Cone(W ) ∩
{u = 1}. The map h 7→ h  W1 is a representation ρ1 : H → C(W1 )≥0 , and the
projection w 7→ w/u(w) from W to W1 is the homeomorphism R of Lemma 2.1(iii).
Lemma 2.3. Let H, u, ρ1 be as in Definition 2.2. Then the homeomorphism F :
Eu → W1 of Lemma 2.1(ii) amounts to F (e) = w iff e(h) = h(w) for every h ∈ H.
Moreover, given any integer b ≥ 1, F induces a bijection between:
(a) the discrete states e ∈ Eu of denominator b;
(b) the rational points w ∈ W1 of denominator b;
(c) the primitive points in Cone(W ) ∩ {u = b}.
In particular, each of the above three sets is finite.
Proof. Since u  W1 = 1l, the statement about F is obvious. Let h1 , . . . , hn , hn+1
be the ρ-counterimages of the generators x1  W , . . ., xn  W , 1l  W of M (W )≥0 .
Then H is generated by h1 , . . . , hn+1 , and hi = xi , for i = 1, . . . , n + 1. Let now
6 G. PANTI, D. RAVOTTI

F (e) = w = (α1 · · · αn+1 )tr ∈ W1 . Then e(hi ) = αi , and e[H] is the intersection
of R≥0 with the subgroup of R generated by α1 , . . . , αn+1 . Therefore e is discrete
of denominator b iff the subgroup of R generated by α1 , . . . , αn+1 is b−1 · Z iff
bα1 , . . . , bαn+1 ∈ Z and the group they generate is all of Z iff bα1 , . . . , bαn+1 are
relatively prime integers iff w is a rational point of denominator b. This establishes
the bijection between (a) and (b). To every rational point w ∈ W1 there  corresponds
the unique primitive point den(w)w ∈ Cone(W ), and u den(w)w = den(w) ·
u(w) = den(w), so (b) and (c) are in bijection. 
We need some further tools from piecewise-linear topology: see [29], [34], [12]
for full details. A rational polytope S is the convex hull of finitely many points
of Qn+1 ; its (affine) dimension is the maximum integer d = dim(S) such that S
contains d + 1 affinely independent points. The affine subspace A = aff(S) of Rn+1
spanned by S is then d-dimensional and defined over Q (i.e., it is the 0-set of finitely
many affine polynomials with rational coefficients). A face of S is S itself or the
intersection of S with an hyperplane π such that S is entirely contained in one of
the two closed halfspaces determined by π; the empty set is a face, and every face
different from S is proper. The index of S is the least integer b ≥ 1 such that
bA ∩ Zn+1 6= ∅ [3, p. 518], and the relative interior of S, relint(S), is the topological
interior of S in A or, equivalently, the set of all points of S which do not lie in a
proper face.
Definition 2.4. Let S, A, b be as above. If d = 0, then S = A = {w} for some w ∈
Qn+1 , and we define νA (S) = 1. Assume d > 0. Then there exists a (nonunique)
affine isomorphism ψ : bA → Rd that maps bA ∩ Zn+1 bijectively to Zd . Let us
abuse language by writing b : A → bA for the map w 7→ bw. We define the relative
volume form ΩA on A as
1
ΩA = d+1 · the pull-back via ψ ◦ b of the standard
b
volume form dx̄ = dx1 ∧ · · · ∧ dxd on Rd .


This amounts to saying that, for every continuous function f : A → R with compact
support (more generally, every Riemann-integrable function), we have
Z Z
1
f ΩA = d+1 f ◦ b−1 ◦ ψ −1 dx̄. (1)
A b R d

Up to sign ΩA does not depend


R on the choice of ψ, and we always assume that ψ
has been chosen so that A f ΩA ≥ 0 for f ≥ 0. The measure νA induced by ΩA
on A is then the d-dimensional Lebesgue measure, appropriately normalized.
Given a Riemann-measurable subset E of A, the identity (1) yields the explicit
formula
Z
νA (E) = 1lE ΩA
A
Z
1
= d+1 1lE ◦ b−1 ◦ ψ −1 dx̄
b Rd
Z (2)
1
= d+1 1lψ[bE] dx̄
b Rd
1
= d+1 λd ψ[bE] ,

b
MEASURES INDUCED BY UNITS 7

where λd is the usual d-dimensional Lebesgue measure on Rd .


Example 2.5. For the reader’s convenience we give here a direct construction for
the map ψ of Definition 2.4, and provide an example. Adopting the above notation,
assume that w ∈ bA ∩ Zn+1 . Then bA − w is a d-dimensional linear subspace of
Rn+1 , and is defined over Q (equivalently, over Z). It follows that there exists a
Z-basis m1 , . . . , md , md+1 , . . . , mn+1 of Zn+1 whose first d elements constitute a
Z-basis for (bA − w) ∩ Zn+1 . Let e1 , . . . , ed denote the standard basis of Rd and let
ϕ : Rn+1 → Rd denote the unique linear map sending mi to ei if i ≤ d, and to 0
otherwise. Then the map ψ(v) = ϕ(v − w) is an affine isomorphism as described in
Definition 2.4. Any other isomorphism ψ 0 : bA → Rd sharing the same properties
must be of the form ψ 0 = t ◦ g ◦ ψ, where g is a linear automorphism of Rd induced
by a matrix M in the group GLd Z of all invertible d × d matrices with integer
entries (this corresponds to choosing a basis for (bA − w) ∩ Zn+1 different from
m1 , . . . , md ), and t is the translation by some vector in Zd ⊂ Rd (this corresponds
to choosing an element of bA ∩ Zn+1 different from w). By multilinear algebra,
changing ψ with ψ 0 merely replaces ΩA with det(M )ΩA and, since det(M ) = ±1,
up to sign ΩA does not depend on ψ.
As an example, consider the following five simplexes in R3 :
• S1 = the convex hull of (0, 1/3, 1), (1/3, 1, 1), (1/9, 8/9, 1);
• S2 = the convex hull of (1/2, 1/4, 1), (1, 1/2, 1);
• S3 = the convex hull of (1/3, 3/5, 1), (1/3, 1, 1);
• S4 = {(1/2, 1/2, 1)};
• S5 = {(2/7, 1/7, 1)}.
All of S1 , . . . , S5 lie in the unit square of the hyperplane {x3 = 1}; we provide a
sketch for the reader’s convenience

S3

S1
S4

S2

S5

Let λ2 , λ1 , λ0 be the usual 2-dimensional, 1-dimensional, 0-dimensional


√ Lebesgue
measures on {x3 = 1}; we have λ2 (S1 ) = 1/18, λ1 (S2 ) = 5/4, λ1 (S3 ) = 2/5,
λ0 (S4 ) = λ0 (S5 ) = 1.
Now, the affine span of S1 is all of {x3 = 1}, which intersects Z3 nontrivially;
hence the index of S1 is 1 and νaff(S1 ) = λ2 . The affine span of S2 again contains
points in Z3 , e.g., w = (0, 0, 1). An appropriate affine isomorphism ψ : aff(S2 ) → R
(i.e., one that establishes a bijection between aff(S2 ) ∩ Z3 and Z ⊂ R) is determined
8 G. PANTI, D. RAVOTTI

by mapping w to 0 and v = (2, 1, 1) to 1. The line segment [w, v] has thus νaff(S2 ) -
√ √
measure 1 and λ1 -measure 5, so that νaff(S2 ) = ( 5)−1 λ1 .
The affine span of S3 does not contain points in Z3 , and neither does 2 aff(S3 ). On
the other hand, 3 aff(S3 ) is the line passing through w = (1, 0, 3) and v = (1, 1, 3),
so the index of S3 is 3; we can take ψ : 3 aff(S3 ) → R as the unique affine map that
sends w to 0 and v to 1. By (2) we have
length of ψ[3S3 ]
νaff(S3 ) (S3 ) = .
9
One easily checks that ψ[3S3 ] is the interval [9/5, 3] in R and concludes that
νaff(S3 ) (S3 ) = 2/15; since λ1 (S3 ) = 2/5, we have νaff(S3 ) = 3−1 λ1 .
Finally, νaff(S4 ) = νaff(S5 ) = λ0 by definition.
If d > 0 and b = 1 then νaff(S) (S) is the relative volume of S [2, §5.4]. More
generally we have the following lemma.
Lemma 2.6. Let d > 0 and let S be a d-dimensional simplex, not necessarily
rational, whose vertices w1 , . . . , wd+1 lie on {a = 1}, for some a ∈ Hom(Zn+1 , Z).
Assume that the subspace V spanned by w1 , . . . , wd+1 in Rn+1 is defined over Q,
and let m1 , . . . , md+1 be a Z-basis for the free Z-module Zn+1 ∩ V . Let M ∈
GLd+1 R be defined by (w1 · · · wd+1 ) = (m1 · · · md+1 )M . Then for every affine
function f : S → R we have

|det(M )| f (w1 ) + · · · + f (wd+1 )
Z
f dνaff(S) = .
S (d + 1)!
R
Proof. It is obvious that S f dλ, where λ is any multiple of the d-dimensional
Lebesgue measure, is the product of λ(S) and the average value of f over the
vertices of S. Since νaff(S) is such a multiple, we just need to check the stated
formula for f = 1l, namely
|det(M )|
νaff(S) (S) = . (3)
d!
Writing b for the least positive integer such that b aff(S) contains an integer point,
we have by definition νaff(S) (S) = b−(d+1) νaff(bS) (bS), so everything boils down to
proving
|det(bM )|
νaff(bS) (bS) = . (4)
d!
Now, the strip V ∩ {0 < a < 1} does not contain integer points and hence nei-
ther does the strip V ∩ {0 < a < b}. It follows that Zn+1 ∩ V has a Z-basis
(r1 , . . . , rd+1 ) with r1 , . . . , rd ∈ V ∩ {a = 0} and rd+1 ∈ V ∩ {a = b} = aff(bS). Let
(bw1 · · · bwd+1 ) = (r1 · · · rd+1 )R; then R ∈ GLd+1 R has last row (1 · · · 1).
As in Example 2.5 we define ψ : aff(bS) → Rd by sending rd+1 to 0 and ri + rd+1
to the i-th element ei of the standard basis of Rd . Thus νaff(bS) (bS) is the ordinary
volume of ψ(bS) in Rd , namely |det(P )|/d!, where P is the (d + 1) × (d + 1) matrix
(see, e.g., [32]) whose last row is (1 · · · 1) and whose upper d × (d + 1) minor P 0 is
defined by
(ψbw1 · · · ψbwd+1 ) = (e1 · · · ed )P 0 .
We claim that P = R. Indeed, let U be the (d + 1) × (d + 1) matrix that has 1
along the main diagonal and the last row, and 0 otherwise. Then
(bw1 · · · bwd bwd+1 ) = (r1 + rd+1 · · · rd + rd+1 rd+1 )U −1 R,
MEASURES INDUCED BY UNITS 9

and each column of U −1 R adds up to 1 (because (1 · · · 1 1)U −1 = (0 · · · 0 1), and R


has last row (1 · · · 1 1)). Hence we get
(ψbw1 · · · ψbwd ψbwd+1 ) = (e1 · · · ed 0)U −1 R = (e1 · · · ed 0)R.
Therefore the d × (d + 1) upper minor of R is P 0 , and since the last row of R is
(1 · · · 1) we have P = R, as claimed.
We have thus proved (4) for a specific choice —namely (m1 · · · md+1 ) =
(r1 · · · rd+1 ), whence bM = R— of the basis of Zn+1 ∩ V . But any other choice is
the image of (r1 · · · rd+1 ) by a matrix in GLd+1 Z, and hence (4) remains valid. 
Example 2.7. Let S1 , S2 , S3 be as in Example 2.5; we can take a = x3 . The
subspace V1 spanned by the vertices of S1 is all of R3 , so we can take the standard
basis of R3 as m1 , m2 , m3 , whence
 
0 1/3 1/9
M1 = 1/3 1 8/9 .
1 1 1
By (3) we have νaff(S1 ) (S1 ) = |det(M1 )|/2! = 1/18, in agreement with the direct
checking of Example 2.5.
The subspace V2 spanned by the vertices of S2 is {x1 − 2x2 = 0}, that intersects
Z3 is the free Z-module Z(0, 0, 1) + Z(2, 1, 0). The matrix M2 is then determined
by    
1/2 1 0 2
1/4 1/2 = 0 1 M2 ,
1 1 1 0
hence  
1 1
M2 = ,
1/4 1/2
and νaff(S2 ) (S2 ) = |det(M2 )|/1! = 1/4. The computation for S3 is analogous.
Definition 2.8. A d-dimensional rational polytopal complex is a finite set Σ of
rational polytopes in Rn+1 such that each face of each element of Σ belongs to Σ,
every two elements intersect in a —possibly empty— common face, at least one
element is d-dimensional and none is l-dimensional, for d < l ≤ n + 1. If all the
elements of Σ are simplexes, we say that Σ is a simplicial complex. For 0 ≤ l ≤ d,
let Σmax (l) be the set of all l-dimensional polytopes of Σ which are not properly
contained
S in any element of Σ. A rational polytopal set W is the underlying set
W = Σ of some rational polytopal complex Σ.
S S
Lemma 2.9. Let W = Σ = Π be a polytopal set. Then:
(i) If S ∈ Σ is l-dimensional, then ΠS = {F ∩ P : F is a face of S and P ∈ Π}
is a pure polytopal complex (i.e., (ΠS )max (r) = ∅ for every r < l).
(ii) If S ∈ Σmax (l), then there exists P ∈ Πmax (l) such that S ∩ P is l-
dimensional.
(iii) Σmax (l) = Πmax (l), for every l.
S S
(iv) Σ and Π have the same dimension.
Proof. (i) Clearly ΠS is a polytopal complex [29, 2.8.6]. Let R ∈ (ΠS )max (r) for
some r ≤ l, and let w be a point in relint(R). Then w ∈/ R0 for any R 6= R0 ∈ ΠS ,
n+1
and therefore Sthere exists an open ball B ⊂ R centered at w and such that
B ∩ R = B ∩ ΠS = B ∩ S. The latter set is l-dimensional, and hence r = l.
10 G. PANTI, D. RAVOTTI

(ii) Let R ∈ (ΠS )max (l) be as in (i). Then R = S ∩P for some P ∈ Πmax (p), with
l ≤ p. By (i), ΣP is a pure p-dimensional complex, and hence S ∩ P is contained
in some p-dimensional element S 0 ∩ P of ΣP . Therefore (S ∩ S 0 ) ∩ P = S ∩ P is
p-dimensional and l ≥ dim(S ∩ S 0 ) ≥Sp; thus S P ∈ Πmax (l).
max
(iii) Let S ∈ Σ (l). Then S = ΠS = (ΠS )max (l), and by the proof of (ii)
max
every R ∈ (ΠS ) (l) is contained in some P ∈ Πmax (l). This shows the left-to-right
inclusion, and the other inclusion is analogous.
(iv) is immediate from (iii). 

Definition 2.10. The dimension of the polytopal set W is the dimension of any
polytopal complex Σ of which W is the underlying set; this makes sense because
of Lemma 2.9(iv). Let 0 ≤ l ≤ d = dim(W ). The l-dimensional support of W
is the set of all hyperplanes of the form aff(S), for some S ∈ Σmax (l). The set
A1 , A2 , . . . , Arl of l-dimensional supporting hyperplanes is finite —possibly empty—
and, by Lemma 2.9(ii), depends on W only. For every l such that W has at least
one l-dimensional supporting hyperplane, and for every Borel subset B of W , define
l
(B) = νA1 A1 ∩ B ∩ Σmax (l) + · · · + νArl Arl ∩ B ∩ Σmax (l) .
S  S 
νW
l
Since νAi (Ai ∩ Aj ) = νAj (Ai ∩ Aj ) = 0 for i 6= j, one checks easily that νW is a
Borel finite measure on W .
0
Note that νW is the counting measure on the set of isolated points of W . As we
l
will be concerned with the measures νW , for l < d, only in the last section of this
d
paper, we save notation by writing νW for νW .

3. Asymptotic distribution of primitive points


From Lemma 2.3 it is clear that in order to deal with Cesàro means of discrete
states we need some understanding of the distribution of primitive points in rational
cones. For d ≥ 1 Jordan’s generalized totient ϕd : N → N is the arithmetical
function defined by
X k Y  1

d d
ϕd (k) = µ h =k 1− d ,
h p
h|k p|k
p prime

where µ is the Möbius function, defined by µ(k) = 0 if k is not squarefree, µ(k) = 1


if k is the product of an even number of distinct prime factors, and µ(k) = −1
otherwise. Clearly ϕ1 is Euler’s totient. We have
X 1
Φd (k) = ϕd (t) = k d+1 + O(k d ); (5)
(d + 1)ζ(d + 1)
t≤k

here the first identity is a definition while the second one, which involves Riemann’s
zeta function ζ, is well known [25, pp. 193–195]. 
As usual, the meaning of identities α(k) = β(k) + O γ(k) such as (5) above is
that there exists a constant C > 0satisfying |α(k) − β(k)| < Cγ(k) for every k.
Analogously, α(k) = β(k) + o γ(k) means that limk→∞ α(k) − β(k) /γ(k) = 0.
We will need the following fact about Cesàro convergence.
Lemma 3.1. Let α : N → R be bounded, let 0 = n0 < n1 < n2 < · · · be a strictly
increasing sequence of natural numbers, and let β ∈ R. Then:
MEASURES INDUCED BY UNITS 11

(i) if
1 X
lim α(t) = β,
k→∞ nk − nk−1
nk−1 <t≤nk

then
1 X
lim α(t) = β; (6)
k→∞ nk
t≤nk

(ii) if (6) holds and limk→∞ (nk − nk−1 )/nk = 0, then


1X
lim α(t) = β.
k→∞ k
t≤k

Proof. (i) is due to Cauchy [7, p. 378], and (ii) is [19, Lemma 2.4.1]. 

Lemma 3.2. Let S ⊆ [0, 1]n be a rational polytope of dimension d ≥ 1. Let


u = a1 x1 + · · · + an xn + an+1 be an affine function with integer coefficients which
is strictly positive on S, and let u be its homogeneous correspondent. Let c be a
positive multiple of the index of the d-dimensional rational polytope S1 = Cone(S)∩
{u = 1}, and let Ξ(S, t) be the number of primitive points in Cone(S) ∩ {u ≤ t}.
Then
Ξ(S, ck)
lim = νaff(S1 ) (S1 ).
k→∞ Φd (ck)

S1 , the integer points in Cone(S) ∩ {u ≤


Proof. By the definition of the index b of S
ck} are contained in the disjoint union {tbS1 : 1 ≤ t ≤ ck/b}. Let LbS1 (t)
(respectively, PbS1 (t)) P
be the number of integer (respectively, primitive) points in
tbS1 . Since LbS1 (t) = h|t PbS1 (h), we have by Möbius inversion
X t
PbS1 (t) = µ LbS1 (h).
h
h|t

By the Ehrhart theory [31, §4.6.2], [2, Theorem 3.23], LbS1 (h) is a quasipolynomial
of degree d and finite period e, with e dividing the lcm of the denominators of the
vertices of bS1 . More precisely, LbS1 (h) has the form
LbS1 (h) = νaff(bS1 ) (bS1 )hd + ld−1 (h)hd−1 + · · · + l1 (h)h + l0 (h),
with ld−1 (h), . . . , l0 (h) rational numbers that only depend on the residue class of h
modulo e. Fix a number M such that |lj (h)| ≤ M for every 0 ≤ j < d and 1 ≤ h.
We thus obtain
X  t  X 
PbS1 (t) = µ νaff(bS1 ) (bS1 )hd + lj (h)hj
h j
h|t
XX  t  
j
= νaff(bS1 ) (bS1 )ϕd (t) + µ lj (h)h
j
h
h|t

= νaff(bS1 ) (bS1 )ϕd (t) + o ϕd (t) ,
because      
1 X X t j dM X t d−1
µ lj (h)h ≤ µ h ,
ϕd (t) j h ϕd (t) h

h|t h|t
12 G. PANTI, D. RAVOTTI

and the latter goes to 0 as t goes to infinity [28, §2]. Therefore


X
Ξ(S, ck) = PbS1 (t)
t≤ck/b
X 
= νaff(bS1 ) (bS1 )Φd (ck/b) + o ϕd (t) .
t≤ck/b

So we get
P  P 
Ξ(S, ck) Φd (ck/b) t≤ck/b o ϕd (t) t≤ck/b o ϕd (t)

− νaff(bS1 ) (bS1 ) = ≤ P ,

Φd (ck) Φd (ck) Φd (ck) t≤ck/b ϕd (t)


and the last expression goes to 0 as k goes to infinity by Lemma 3.1(i). Finally,
due to the asymptotic estimate (5),
Φd (ck/b) 1
lim νaff(bS1 ) (bS1 ) = νaff(bS1 ) (bS1 ) d+1 = νaff(S1 ) (S1 ).
k→∞ Φd (ck) b


4. The main result


Lemma 4.1. Let W ⊆ Rn+1 be a rational polytopal set, and let ȳ = y1 , y2 , y3 , . . .,
z̄ = z1 , z2 , z3 , . . . be sequences of points in W . Suppose that for every rational
polytope T ⊆ W the Cesàro average of 1lT along ȳ exists, say
1X
lim 1lT (yt ) = A(1lT , ȳ).
k→∞ k
t≤k

Then:
(i) for every f ∈ C(W ), the Cesàro average A(f, ȳ) exists;
(ii) the functional A(−, ȳ) : C(W ) → R is linear, positive (i.e., A(f, ȳ) ≥ 0 if
f ≥ 0), and normalized (i.e., A(1lW , ȳ) = 1). By the Riesz representation
theorem, A(−, ȳ) is induced by integration w.r.t. a uniquely defined Borel
probability measure on W ;
(iii) if, for every T , the average A(1lT , z̄) exists and equals A(1lT , ȳ), then
A(f, z̄) = A(f, ȳ) for every f ∈ C(W ).
Proof. Since by hypothesis the Cesàro averages along ȳ exist for all characteristic
functions 1lT , they exist for all step functions, i.e., all R-linear combinations of such
characteristic functions. The proof is now straightforward from the fact that the
set of step functions is dense in L∞ (W ) (the Banach space of all bounded functions
on W with the topology of uniform convergence). 
The following is our main result.
Theorem 4.2. Let H be a finitely presented cancellative hoop in which a unit u
has been fixed. Let e1 , e2 , e3 , . . . be an enumeration without repetitions of all the
discrete extremal states of (H, u) according to nondecreasing denominators. Then
for every h ∈ H the limit
1X
mu (h) = lim et (h)
k→∞ k
t≤k

exists, does not depend on the enumeration, and the function mu : H → R≥0 thus
defined is an automorphism-invariant state.
MEASURES INDUCED BY UNITS 13

Proof. Let n ≥ 0 be the least integer such that H can be generated by n+1 elements
h1 , . . . , hn+1 .
Remark 4.3. We do not require h1 , . . . , hn+1 ≤ u. In particular, the MV-algebra
Γ(H, u) is still finitely generated —a simple application of the Riesz decomposition
property [14, Proposition 2.2]—, but the least cardinality of a generating set may be
larger than n + 1. For example, let H = Free2 (CH), and let u = x1 ∨ (−kx1 + 1) for
some integer k ≥ 1. Then n = 1, because H is not generable by a single element.
On the other hand, Eu contains k + 2 discrete states of denominator 1, namely
the states h 7→ h(w)/u(w) for w ∈ {0, 1, 1/2, 1/3, 1/4, . . . , 1/(k + 1)}. The usual
representation of Γ(H, u) as an MV-algebra, i.e., as a quotient of Freem (MV) for
some integer m, assumes that u is represented by the constant function 1l. Now,
the only points in the unit cube [0, 1]m that correspond to states of denominator 1
w.r.t. the unit 1l are the 2m vertices of the cube. In our case we need k + 2 such
points, hence presenting Γ(H, u) as an MV-algebra requires at least log2 (k + 2)
generators.
Our H can then be presented as hx1 , . . . , xn+1 ; f1 (x̄) = g1 (x̄), . . . , fr (x̄) = gr (x̄)i.
As cited in §2, Freen+1 (CH) ' M ([0, 1]n )≥0 . Also, W = {w ∈ [0, 1]n : fi (w, 1 −
(w1 ∨ · · · ∨ wn )) = gi (w, 1 − (w1 ∨ · · · ∨ wn )) for every i} is a rational polytopal
set of dimension 0 ≤ d ≤ n, and the hoop homomorphism ρ : H → M (W )≥0
defined by ρh1 = x1  W , . . ., ρhn = xn  W , ρhn+1 = 1l − (x1 ∨ · · · ∨ xn )  W is
a McNaughton representation. We adopt the notation of Definition 2.2.
The case d = 0 is trivial. Indeed, in this case W is just a finite set of rational
points, and by Lemma 2.1(ii) the given enumeration is finite, say e1 , e2 , . . . , er .
Therefore, mu is always defined and is a state. Given any automorphism σ of H
such that σ(u) = u, the sets {e  1 , e2 , . . . , er } and {e1 ◦ σ, e2 ◦ σ, . . . , er ◦ σ} are equal,
and hence mu (h) = mu σ(h) for every h.
Let now d ≥ 1; by Lemma 2.3, the enumeration e1 , e2 , e3 , . . . corresponds to
an enumeration w1 , w2 , w3 , . . . of all primitive points in Cone(W ) according to
nondecreasing values of u. Let wt0 = wt /u(wt ) be the point of intersection of W1
with the ray R≥0 wt ; by Lemma 2.3 et (h) = h(wt0 ), so we must show that
1X
mu (h) = lim h(wt0 ) (7)
k→∞ k
t≤k

exists.
Since h  W1 is a continuous function, we can use Lemma 4.1(i) to prove the
convergence of (7). Let then T be a rational polytope contained in W1 . S Us-
ing [29, 2.8(6)] we can construct a polytopal complex Σ such that W1 = Σ
and T is a union of elements of Σ; we can thus safely assume T ∈ Σ. Note
that we do not need that Σ is a simplicial complex, nor that it satisfies any
unimodularity condition (see [23] for unimodular simplicial complexes). Let c
be the lcm of the d-indices of the d-dimensional polytopes in Σ. The set of
primitive points in Cone(W ) is then partitioned in blocks B1 , B2 , B3 , . . ., where
Bk = {w ∈ Cone(W ) : w is primitive and c(k − 1) < u(w) ≤ ck}. For every k ≥ 1,
let nk be such that wnk ∈ Bk and wnk +1 ∈ Bk+1 . Generalizing the notation in
Lemma 3.2, let Ξ(W, t) be the number of primitive points in Cone(W ) ∩ {u ≤ t}.
Then X
Ξ(W, ck) = {(S) · Ξ(S, ck) : S ∈ Σ},
14 G. PANTI, D. RAVOTTI

where (S) is either +1, 0 or −1, according to the inclusion-exclusion principle


applied to S w.r.t. the combinatorial structure
 of Σ. Since (S) is necessarily +1 if
S is d-dimensional, and Ξ(S, ck) = o Φd (ck) if dim(S) < d, from Lemma 3.2 and
the estimate (5) we obtain
νW1 (W1 )
k d+1 + o(k d+1 ).

Ξ(W, ck) = νW1 (W1 )Φd (ck) + o Φd (ck) = (8)
(d + 1)ζ(d + 1)
We thus have, for every k ≥ 1,
1 X 1 X Ξ(T, ck) Ξ(T, ck)/Φd (ck)
1lT (wt0 ) = 1lCone(T ) (wt ) = = .
nk nk Ξ(W, ck) Ξ(W, ck)/Φd (ck)
t≤nk t≤nk

Therefore, by Lemma 3.2, (5) and (8), we obtain


1 X νW1 (T )
lim 1lT (wt0 ) = ; (9)
k→∞ nk νW1 (W1 )
t≤nk

note that νW1 (T ) = 0 if dim(T ) < d. From (8) it is immediate that limk→∞ (nk −
nk−1 )/nk = 0, so we apply Lemma 3.1(ii) and conclude
1X νW1 (T )
lim 1lT (wt0 ) = . (10)
k→∞ k νW1 (W1 )
t≤k

Any enumeration of the discrete extremal states of (H, u) according to nonde-


creasing denominators must induce the same partition B1 , B2 , B3 , . . . of the prim-
itive points in Cone(W ). Since the limits (9) and (10) only depend on this latter
block partition, Lemma 4.1(iii) shows that (7) does not depend on the enumeration.
By Lemma 4.1(ii), mu is a state on (H, u).
Finally, let σ be an automorphism of H that fixes u. As in the d = 0 case,
precomposition with σ is a homeomorphism of Eu preserving the discrete states
and their denominators. Therefore e1 ◦ σ, e2 ◦ σ, e3 ◦ σ, . . . is another enumeration
of the discrete extremal states of (H, u) according to nondecreasing denominators.
By the above
 remarks, the Cesàro limit of h w.r.t. the new enumeration, namely
mu σ(h) , equals the Cesàro limit of h w.r.t. the old enumeration, namely mu (h).
Hence mu is automorphism-invariant. 

Corollary 4.4. Adopt the hypotheses of Theorem 4.2 and the relative notation.
Then, for every h ∈ H, we have
Z
1
mu (h) = h  W1 dνW1 .
νW1 (W1 ) W1
Proof. If d = 0, then νW1 is the counting measure on the finite set W1 =
{w10 , . . . , wr0 }, which is in bijection with {e1 , . . . , er }; since et (h) = h(wt0 ), our
statement is immediate. If d ≥ 1 then, by Lemma 4.1(ii) applied to W1 , mu (h) is
given by integration of h  W1 w.r.t. a certain Borel finite measure on W1 . By (10),
that measure is νW1 /νW1 (W1 ). 

By realizing W1 as the underlying set of a simplicial complex ∆ such that h  ∆


is affine for every S ∈ ∆ (this is always possible, see, e.g., the proof of [23, The-
orem 4.1]), Lemma 2.6 provides an efficient way for computing the expression for
mu (h) given by Corollary 4.4; we will make use of this in §7.
MEASURES INDUCED BY UNITS 15

5. Topological dimension without topology


The maximal spectrum of H is an invariant of H, and in particular its topological
dimension —which is an integer ≥ 0 [16, Definition III 1], [10, Definition 1.1.1]—
is implicitly determined by H. The results in §3 and §4 provide an expression for
this integer that bypasses topology; we will return to dimensional issues in the final
section of this paper.
Theorem 5.1. Let H be a finitely presented cancellative hoop. Let u ∈ H be a
unit and let ](Eu ≤ t) denote the number of discrete extremal states of (H, u) of
denominator ≤ t. Then the limit

log ](Eu ≤ t)
l = lim
t→∞ log(t)
exists and is an integer not depending on u. We have l = 0 iff Eu is finite iff
MaxSpec H has topological dimension 0. Otherwise, 2 ≤ l ≤ (the minimum cardi-
nality of a generating set for H), and MaxSpec H has topological dimension l − 1.
Proof. As in the proof of Theorem 4.2, we identify H with M (W )≥0 , where W
is a polytopal set in [0, 1]n and n + 1 is the minimum cardinality of a generating
set. Clearly l = 0 if Eu is a finite set. We know from Lemma 2.1(ii) that W ,
MaxSpec H and Eu are canonically homeomorphic, so if Eu is not a finite set then
W has dimension d ≥ 1. Let  then c be as in the proof of Theorem 4.2 and let
k(t) = dt/ce. Then c k(t) − 1 < t ≤ ck(t) and
  
log Ξ W, c(k(t) − 1) log Ξ W, t) log Ξ W, ck(t)
 ≤ ≤ .
log ck(t) log(t) log c(k(t) − 1)
Since ](Eu ≤ t) = Ξ(W, t), our statements follow readily from (8) and the highly
nontrivial fact that the dimension of W as a polytopal set coincides with its topo-
logical dimension [10, pp. 101–102]. 
Remark 5.2. In general, in the statements of Theorem 4.2 and of Theorem 5.1 it is
not possible to replace “finitely presented” with “finitely generated”. For example,
let 0 < α < 1 be any irrational number, and let W ⊂ [0, 1]2 be the line segment
of extrema (α, 0) and (α, 1). Then M (W )≥0 , 1l is a three-generated, countably
presented cancellative hoop with unit that has no discrete extremal state at all.

6. Absolute continuity
Let H be a separating subhoop of some C(X)≥0 (i.e., the inclusion map of H
in C(X)≥0 is a representation). Assume that H, as an abstract hoop, is finitely
presentable. Let u1 , u2 ∈ H be two arbitrary units; we are not assuming any
structure either on X (besides being a compact Hausdorff space), or on the elements
of H (besides being continuous functions). Note that u1 and u2 may lie in different
orbits under the action of the automorphism group of H; equivalently, the MV-
algebras Γ(H, u1 ) and Γ(H, u2 ) may not be isomorphic. This is the case, e.g., of
the functions 1l and u of Remark 4.3, which are both units in M ([0, 1])≥0 .
Let m1 , m2 be the automorphism-invariant states determined by Theorem 4.2.
By Lemma 2.1(iv) there exist two uniquely determined Borel finite measures µ1 , µ1
on X such that Z
mi (h) = h dµi ,
X
16 G. PANTI, D. RAVOTTI

for every h ∈ H and i = 1, 2.


Theorem 6.1. Under the above hypotheses and notation, each of µ1 and µ2 is
absolutely continuous w.r.t. the other. More precisely, let d be the topological
dimension of X, as given by Theorem 5.1. We have:
(i) if d = 0, then
u1
dµ2 = dµ1 ; (11)
u2
(ii) if d ≥ 1, then
 d+2
u1
dµ2 = C dµ1 ,
u2
where C is the constant
−1
ud+2
Z
1
C= d+1
dµ1 .
X u2

We devote the rest of this section to the proof of Theorem 6.1. The case d = 0 is
 X = {p1 , · · · , pr }, in bijection with
trivial. Indeed, in this case X is a finite set, say
Eui , for i = 1, 2, via pj 7→ h 7→ h(pj )/ui (pj ) . All extremal states are discrete, so
mi (h) = r−1 1≤j≤r h(pj )/ui (pj ). The measure µi is therefore
P

1 X 1
µi = δp ,
r ui (pj ) j
1≤j≤r

where δpj is the Dirac unit measure at pj , and the formula (11) is clear.
Assume now d ≥ 1; we shall prove that there exists a constant C such that
Z Z  d+2
u1
f dµ2 = C f· dµ1 , (12)
X X u2
for every continuous function f : X → R. As in the proof of Theorem 4.2,
we construct a McNaughton representation ρ : H → M (W )≥0 for a certain d-
dimensional polytopal set W ⊆ [0, 1]n . We adopt the notation in the proof of The-
orem 4.2; in particular, for i = 1, 2, ui is the homogeneous correspondent of ρui , and
Wi = Cone(W ) ∩ {ui = 1}. We thus have two representations ρi : H → C(Wi )≥0 ,
given by ρi h = h  Wi . We construct a rational polytopal complex Σ having W as
its underlying set and such that both ρu1 and ρu2 are affine on each polytope of Σ.
ThenSΣi = {Si : Si = Cone(S) ∩ Wi , for S ∈ Σ} is a rational polytopal complex
and Σi = Wi . Let S ∈ Σ be d-dimensional, let bi be the index of Si , and choose
affine isomorphisms ψi as in Definition 2.4 from aff(bi Si ) to Rd . We display our
data in a diagram

b2 S o b2
SO 2 / W2
O 2 O
ψ2 R2
G P P
}
Rd o b1 S 1 o S1 / W1 /X
ψ1 b1 R1

The unnamed maps are inclusions, and P, R1 , R2 are the homeomorphisms of


Lemma 2.1(iii) induced by ρ1 , ρ2 , and the identity representation; as in Defini-
tion 2.4 bi denotes multiplication by bi . We define G = b2 ◦ P ◦ b−1
1 ; by explicit
computation G(w) = b2 w/u2 (w). Everything is commutative, except for the left-
hand-side triangle, which in general is not. Let then F = ψ2 ◦ G ◦ ψ1−1 ; it is a
MEASURES INDUCED BY UNITS 17

diffeomorphism between the interiors of the d-dimensional polytopes ψ1 b1 S1 and


ψ2 b2 S2 .

Lemma 6.2. Let p be a point in the interior of ψ1 b1 S1 ; then the absolute value
jF (p) of the determinant of the Jacobian matrix of F at p has value
 d+1
b2
jF (p) = .
u2 (ψ1−1 p)
Proof. Let T ⊆ ψ1 b1 S1 be a d-dimensional simplex, not necessarily rational,
containing p in its interior. By [30, Theorem 7.24], jF (p) is the limit, for
T shrinking to p, of the ratio vol(F T )/ vol(T ). By definition, that ratio is
νaff(b2 S2 ) (Gψ1−1 T )/νaff(b1 S1 ) (ψ −1 T ). Let V be the (d + 1)-dimensional real vec-
tor space spanned by S inside Rn+1 . Since V is defined over Q, it intersects Zn+1
in a free Z-module of rank d + 1; let m1 , . . . , md+1 be a Z-basis for this module.
Let w1 , . . . , wd+1 be the vertices of ψ1−1 T and let (w1 · · · wd+1 ) = (m1 · · · md+1 )M
for a certain M ∈ GLd+1 R. Then νaff(b1 S1 ) (ψ −1 T ) = |det M |/d! by Lemma 2.6.
Since the vertices of Gψ1−1 T are b2 w1 /u2 (w1 ), . . . , b2 wd+1 /u2 (wd+1 ), and the
argument above applies to νaff(b2 S2 ) as well, one immediately computes

νaff(b2 S2 ) (Gψ1−1 T ) bd+1


2
= .
νaff(b1 S1 ) (ψ −1 T ) u2 (w1 ) · · · u2 (wd+1 )

Letting T shrink to p, the w’s converge to ψ1−1 p and our claim follows. 

Recall that if L : Y → Z is a continuous map between compact Hausdorff spaces


and µ is a Borel measure on Y , then the push-forward of µ by L is the Borel
measure L∗ µ on Z defined by (L∗ µ)(A) = µ(L−1 A). Equivalently,
Z Z
f ◦ L dµ = f dL∗ µ,
Y Z

for every continuous function f : Z → R.


By Lemma 2.1(iii) applied to ρi and the identity representation of H, there exists
a unit fi ∈ C(Wi )≥0 such that h  Wi = fi · (h ◦ Ri ) for every h ∈ H and i = 1, 2.
Taking h = ui we see that
ui  Wi 1
fi = = ◦ Ri ,
ui ◦ Ri ui
and therefore
h
h  Wi = ◦ Ri .
ui
Setting Di = νWi (Wi ), we have from Corollary 4.4
Z Z Z
1 1 h 1 1
mi (h) = h  Wi dνWi = ◦ Ri dνWi = h· dRi∗ νWi ,
Di Wi Di Wi ui Di X ui
and therefore
1
dµi = dRi∗ νWi . (13)
D i ui
18 G. PANTI, D. RAVOTTI

The preliminaries being over, let us fix a continuous function f : X → R. We


choose S ∈ Σmax (d) and compute:
Z Z
f f
◦ R2 dνW2 = ◦ R2 Ωaff(S2 )
S2 u2 S2 u2
Z
1 f
= d+1 ◦ R2 ◦ b−1 −1
2 ◦ ψ2 dx̄
b2 ψ2 b2 S2 u2
Z
1 f
= d+1 ◦ R1 ◦ b−1 1 ◦G
−1
◦ ψ2−1 dx̄
b2 ψ2 b2 S2 u2
Z  
1 f
= d+1 ◦ R1 ◦ b−11 ◦ G −1
◦ ψ −1
2 ◦ F · jF dx̄
b2 ψ1 b1 S1 u2
Z    
f −1 −1 1 −1
= ◦ R1 ◦ b1 ◦ ψ1 · ◦ ψ1 dx̄
ψ1 b1 S1 u2 ud+1
Z   2
f 1
= ◦ R1 ◦ b−11 · d+1 ◦ ψ1−1 dx̄
ψ1 b1 S1 u 2 u2
Z    
f 1
= ◦ R1 · ◦ b 1 ◦ b−1 −1
1 ◦ ψ1 dx̄
ψ1 b1 S1 u2 ud+1
Z  2 
1 f 1
= d+1 ◦ R1 · d+1
◦ b−1 −1
1 ◦ ψ1 dx̄
b1 ψ1 b1 S1 u2 (u 2  W 1 )
Z  
f 1
= ◦ R1 · d+1
Ωaff(S1 )
S u 2 (u 2  W 1)
Z 1 
f 1
= ◦ R1 · d+1
dνW1
S1 u 2 (u 2  W 1)
Z   d+1 
f u1
= · ◦ R1 dνW1 .
S1 u 2 u2
Since νWi is 0 on Si ∩ Ti for every S 6= T ∈ Σmax (d), we obtain from (13)
Z Z
1 f
f dµ2 = dR2∗ νW2
X D 2 X 2u
Z
1 f
= ◦ R2 dνW2
D2 W2 u2
1 X Z f
= ◦ R2 dνW2
D2 max S2 u2
S∈Σ (d)
Z   d+1 
1 X f u1
= · ◦ R1 dνW1
D2 S1 u2 u2
S∈Σmax (d)
Z   d+1 
1 f u1
= · ◦ R1 dνW1
D2 W1 u 2 u2
Z  d+2
1 f u1
= · dR1∗ νW1
D2 X u1 u2
Z  d+2
D1 u1
= f· dµ1 .
D2 X u2
MEASURES INDUCED BY UNITS 19

So (12) is proved with C = D1 /D2 . The expression for C in (ii) follows immediately
by taking f = u2 in (12). The proof of Theorem 6.1 is thus complete.

7. The local dimension at an extreme state


It is apparent from the proof of Theorem 4.2 that the state mu neglects the
behaviour of h ∈ H on the underdimensional parts of MaxSpec H. Indeed, if
h, g ∈ H are such that ρh = ρg on Σmax (d), then by Corollary
S
4.4 mu (h) = mu (g),
irrespective of the behaviour of ρh and ρg on W \ Σmax (d). Since measure-
S
theoretic issues neglect sets of measure 0, this fact is quite natural. However, if one
insists on having states capable of detecting the behaviour of the elements of H on
all parts of the spectrum —the faithful states of [22], [23]— then our construction
can be adapted.
First of all, let us recall the inductive definition of the local dimension of a
topological space at a point [16, Chapter III]:
(1) The only set of dimension −1 is the empty set.
(2) A space has local dimension ≤ n at a point p if p has arbitrarily small
neighborhoods whose boundaries have dimension ≤ n − 1.
(3) A space has dimension ≤ n if it has local dimension ≤ n at each of its
points.
(4) A space has local dimension n at p if it has local dimension ≤ n at p, and
does not have local dimension ≤ n − 1 at p.
(5) A space has dimension n if it has dimension ≤ n and does not have dimen-
sion ≤ n − 1.
(6) A space has dimension ∞ if it does not have dimension ≤ n for any n.
In the proof of Theorem 5.1 we already made use of the fact that the dimension
of a polytopal set coincides with its topological dimension (note that throughout
this paper the unqualified word “dimension” always refers to the affine dimension).
The following lemma provides a local version of that fact.
S
Lemma 7.1. Let W = Σ be a d-dimensional polytopal set. For every 0 ≤ l ≤ d,
let
W l = Σmax (l) \ {S ∈ Σ : dim(S) > l},
S S

Ll = {w ∈ W : the local dimension of W at w is l}.


Then
W l = Ll ,
for every l. In particular, the partition
{W l : W l 6= ∅}
of W does not depend on Σ.
Proof. Since both {W l : W l 6= ∅} and {Ll : Ll 6= ∅} are partitions of W , it clearly
suffices to show that W l ⊆ Ll for every l. Let w ∈ W l . Then w does not belong to
n+1
any element of Σ of dimension > l, and hence there exists an
0 0
S open ball B ⊂ R
centered at w and such that B ∩ W = B ∩ W , where W = {S ∈ Σ : dim(S) ≤ l}
is an l-dimensional polytopal set. Thus the local dimension of W at w agrees with
the local dimension of W 0 at w, which is ≤ l again by [10, pp. 101–102]. We thus
have to show that the local dimension of W 0 at w is not ≤ l − 1. If l = 0 this is
clear, so assume l > 0 and choose S ∈ Σmax (l) such that w ∈ S. Choose a ball B
20 G. PANTI, D. RAVOTTI

as above such that relint(S) is not contained in the closure of B. It is then enough
to show the following:
(A) Let U be an open set in W 0 such that w ∈ U ⊆ B, and write M for the
boundary of U in W 0 . Then M ∩ relint(S) contains a point at which the
local dimension of M is not ≤ l − 2.
Let U be as in (A) and let Ū be its closure in W 0 . Then relint(S) is the disjoint
union of relint(S) ∩ U , relint(S) ∩ M , and relint(S) \ Ū . By construction there exist
points p ∈ relint(S) ∩ U and q ∈ relint(S) \ Ū . It follows that every continuum
(i.e., closed connected set) C ⊆ relint(S) containing p and q must intersect M ,
for otherwise (C ∩ U ) ∪ (C \ Ū ) would be a nontrivial disconnection of C. By
Mazurkiewicz’s Theorem [10, Theorem 1.8.19] applied to the open region relint(S)
inside aff(S) ' Rl , the topological dimension of M ∩ relint(S) is not ≤ l − 2.
Therefore there exists a point as required by (A). 
l l
S νW of Definitionl 2.10. Note that, since νW is sup-
We shallSneed the measures
ported on Σmax (l), and Σmax (l) \ W l is a νW -nullset, we have
Z Z Z
l l l
f dνW = S f dνW = f dνW ,
W Σmax (l) Wl
for every Riemann-integrable function f : W → R.
Assume now the hypotheses and the notation of Theorem 4.2; we thus have a
McNaughton representation ρ : H → M (W )≥0 and a homeomorphism F : Eu →
W1 = Cone(W ) ∩ {u = 1} as in Lemma 2.3. By Lemma 7.1, for every 0 ≤ l ≤ d =
dim(Eu ), F restricts to a bijection between Eul = {e ∈ Eu : the local dimension of
Eu at e is l} and W1l . Let 0 ≤ l1 < l2 < · · · < lq = d be the dimensions at which
Euli 6= ∅. For each 1 ≤ i ≤ q, let ei1 , ei2 , ei3 , . . . be an enumeration without repetitions
of all the discrete states in Euli according to nondecreasing denominators. Choose
a weight vector p = (p1 , . . . , pq ) (i.e., p ∈ Rq≥0 and
P
pi = 1).
Theorem 7.2. For every i = 1, . . . , q and every h ∈ H the limit
1X i
mlui (h) = lim et (h) (14)
k→∞ k
t≤k

exists and does not depend on the enumeration. The function mpu : H → R≥0
defined by X
mpu (h) = pi mlui (h) (15)
i
is an automorphism-invariant state, which is faithful (i.e., h 6= 0 implies mpu (h) 6= 0)
iff pi > 0 for every i.
Proof. Every convex combination of automorphism-invariant states is clearly an
automorphism-invariant state, so it suffices to show that every mlui is such a state.
Fix then 0 ≤ l ≤ d such that Eul 6= ∅, Sand say that W is the underlying set
of the polytopal complex Σ. Let W = Σmax (l); then W is an l-dimensional
rational polytopal set which, by Lemma 2.9(iii), depends on W and l only. The
restriction map ρh 7→ ρh  W is a homomorphism from M (W )≥0 to M (W )≥0 , and
the latter hoop is finitely presented because W is rational. Every element of the
enumeration el1 , el2 , el3 , . . . provided by the hypotheses corresponds to a point of W ,
and can thus be seen as a discrete extremal state of M (W )≥0 , with preservation
of denominators. Moreover, all discrete extremal states of M (W )≥0 appear in the
MEASURES INDUCED BY UNITS 21

enumeration, except those corresponding to points of W whose local dimension as


points of W is > l (for example, for l = 1 in the polytopal set W = S1 ∪ · · · ∪ S5 of
Example 2.5, we have W = S2 ∪ S3 , W 1 = W \ S1 , and the point of intersection of
S3 with S1 is the only such point).
Let W 1 = Cone(W ) ∩ {u = 1}, W1l = Cone(W l ) ∩ {u = 1}, and write for
simplicity’s sake f = h  W1 . Let M > 0 be an upper bound for f , and let
f − , f + be the functions (in general not continuous) which agree with f on W1l
but are identically 0 (respectively, M ) on W1 \ W1l . We construct a denominator-
nondecreasing sequence w1 , w2 , w3 , . . . of all rational points of W 1 , containing a
subsequence wk(1) , wk(2) , wk(3) , . . . such that elt corresponds to wk(t) , for every t
(this just amounts to “reinserting” the missing states in the given enumeration). It
is clear that we have, for every t,
k(t) t k(t)
1 X − 1X l 1 X +
f (wi ) ≤ et (h) ≤ f (wi ). (16)
k(t) i=1 t i=1 k(t) i=1
Now, the reinserted points (i.e., the points in the sequence (wi ) which do not belong
to the subsequence) are all contained in W 1 \ W1l , which is a νW l
1
-nullset. By [19,

Exercise 1.12 p. 179] and Theorem 4.2 applied to M (W )≥0 , u  W , the two
l
−1 R l
extreme terms of (16) converge as t → ∞ to νW 1
(W1 ) f dνW 1
. Therefore the
middle term converges to the same value, as was to be shown. The same argument
as in Theorem 4.2 shows the automorphism-invariance of mlu .
Concerning our last statement, the proof of right-to-left implication is analogous
to the proof of the faithfulness of the state s in [23, Theorem 4.1]. Namely, assume
p1 , . . . , pq > 0 and let 0 6= h ∈ H, with the intent of proving mpu (h) > 0. As noted
after Corollary 4.4 we can realize W1 as the underlying set of a rational simplicial
complex ∆ such that h  S is affine for every S ∈ ∆. Applying Corollary 4.4 and
Lemma 2.6, we obtain
q Z
X 1 li
mpu (h) = pi li h  W1 dνW 1
(17)
i=1
ν W1 (W 1 ) W1

q
!
X pi X
li
= li
νW (S)h(S) , (18)
i=1
νW 1
(W1 ) S∈∆max (l ) 1
i

where h(S) is the arithmetical average of h over the vertices of S. All terms pi ,
li li
νW 1
(W1 ), and νW 1
(S) in (18) are > 0. Since h 6= 0 implies h > 0 on at least one
vertex of at least one S ∈ ∆, we have mpu (h) > 0, as desired.
For the reverse implication, let pi = 0 for some 1 ≤ i ≤ q, and let ∆ be a
rational simplicial complex supported on W1 . Since Euli 6= ∅, ∆max (li ) contains a
simplex, say S. Using elementary algebra we construct a Z-basis m1 , . . . , mli +1 of
Zn+1 ∩ (the R-subspace spanned by S) such that the ray R≥0 mj intersects S in its
relative interior, for every j. Let T be the intersection of S with the cone spanned
positively by m1 , . . . , mli +1 ; let also w = m1 + · · · + mli +1 . Since S ∈ ∆max (li ) and
T is contained in the relative interior of S, the simplex T has empty intersection
with every element of ∆ different from S. It readily follows (compare with Mundici’s
theory of Schauder hats [24, §5.3]) that the function h : Cone(W1 ) → R≥0 defined
by:
• h = 0 outside of Cone(T );
22 G. PANTI, D. RAVOTTI

• h(m1 ) = · · · = h(mli +1 ) = 0;
• h(w) = 1;
• for each 1 ≤ i ≤ li + 1, h is linear on the cone spanned positively by
m1 , . . . , mi−1 , w, mi+1 , . . . , mli +1 ;
is continuous and piecewise-linear with integer coefficients. Let h be the unique
element of H such that h is the homogeneous correspondent of ρh. Then h 6= 0
and, since pi = 0 and h = 0 outside Cone(T ), we have mpu (h) = 0 by (17), so mpu
is not faithful. 

Remark 7.3. In the specific case u = 1l and p = (1/q, . . . , 1/q), the state mp1l
of Theorem 7.2 coincides with the state s of [23, Theorem 4.1], provided that
MaxSpec H does not contain isolated points. Indeed, let h ∈ H = M (W )≥0
for some d-dimensional rational polytopal set W ⊆ [0, 1]n ; since u = 1l we
have W1 = W and h  W1 = h. Choose a unimodular complex ∆ supported
on W (see [23] for all relevant definitions), such that h is affine on S each sim-
plex of ∆. Let T be an l-dimensional unimodular simplex contained in ∆max (l)
(not necessarily belonging to ∆), and let w1 , . . . , wl+1 be the vertices of T . By
the unimodularity assumption (den(w1 )w1 · · · den(wl+1 )wl+1 ) is a Z-basis for
Zn+1 ∩ (the R-subspace spanned by T ). Therefore by Lemma 2.6
l 1
νW (T ) = νaff(T ) (T ) = .
l! den(w1 ) · · · den(wl+1 )
l
Since the unimodular simplexes generate the Borel σ-algebra, νW is the measure λl
of [23, Theorem 2.1]. Therefore the identity (18), which in our specific case reads
q
!
p 1X 1 X
li
m1l (h) = li
νW (S)h(S) ,
q i=1 νW (W ) S∈∆max (l)

agrees with the expression (4) on [23, p. 542] for the state s.
If W contains isolated points, say W 0 = {w1 , . . . , wr } =
6 ∅, then the situation is
slightly different. Indeed, the measure µ0 corresponding to our m01l is
1 X
µ0 = δwj ,
r
1≤j≤r
0
while the measure τ determined by (4) on [23, p. 542] is
 X −1 X
1 1
τ0 = · δw .
den(wj ) den(wj ) j
1≤j≤r 1≤j≤r

In general µ0 and τ 0 are different, but they both are automorphism-invariant, since
every automorphism of (H, 1l) must act on W 0 in a denominator-preserving way.
For example, let again W = S1 ∪· · ·∪S5 be the polytopal set of Example 2.5, S4 =
{v}, S5 = {w}. Then den(v) = 2, den(w) = 7, W 0 = {v, w}; take h = x1  W to
be the projection on the first coordinate. Then
Z
1
h dµ0 = h(v) + h(w)

W 2
!
1 1 2 11
= + = ,
2 2 7 28
MEASURES INDUCED BY UNITS 23

while
Z !−1 !
0 1 1 h(v) h(w)
h dτ = + +
W 2 7 2 7
!
14 1 2 19
= + = .
9 4 49 42

The hidden reason for the rigidity of the invariant measure on the higher-
dimensional parts of the spectrum, together with its relative flexibility on the
0-dimensional part, is subtle: the 0-dimensional part is made of finitely many iso-
lated discrete extremal states, while the higher-dimensional parts contain infinitely
many discrete extremal states. Now, in a finite average the input data can be rear-
ranged arbitrarily without affecting the final result, a fact which is definitely false
for infinite averages. Actually, a classical result by Descovich (but dating back to
von Neumann for the real unit interval [19, Theorems II.4.4 and III.2.5]) guarantees
that every regular Borel measure on a compact Hausdorff space with no isolated
points can be realized as in Lemma 4.1(ii) by enumerating appropriately the ele-
ments of any given countable everywhere dense subset. In our case, the set of all
discrete extremal states is such a subset, and in Theorem 4.2 the key constraint on
the resulting measure is the denominator-nondecreasing condition, which becomes
irrelevant precisely on the 0-dimensional part of the spectrum. As that condition
depends crucially on the choice of a unit, it is apparent that units and measures
are inextricably intertwined.

References
[1] E. Badouel, J. Chenou, and G. Guillou. An axiomatization of the token game based on Petri
algebras. Fund. Inform., 77(3):187–215, 2007.
[2] M. Beck and S. Robins. Computing the continuous discretely. Undergraduate Texts in Math-
ematics. Springer, 2007.
[3] M. Beck, S. V. Sam, and K. M. Woods. Maximal periods of (Ehrhart) quasi-polynomials. J.
Combin. Theory Ser. A, 115(3):517–525, 2008.
[4] W. M. Beynon. Duality theorems for finitely generated vector lattices. Proc. London Math.
Soc., 31(3):114–128, 1975.
[5] A. Bigard, K. Keimel, and S. Wolfenstein. Groupes et anneaux réticulés, volume 608 of
Lecture Notes in Math. Springer, 1977.
[6] W. J. Blok and I. M. A. Ferreirim. On the structure of hoops. Algebra Universalis, 43(2-
3):233–257, 2000.
[7] T. J. Bromwich. An introduction to the theory of infinite series. Macmillan and Co., 1908.
Available at the Open Library, http://openlibrary.org/books/OL7073755M.
[8] R. Cignoli, I. D’Ottaviano, and D. Mundici. Algebraic foundations of many-valued reasoning,
volume 7 of Trends in logic. Kluwer, 2000.
[9] A. Dvurečenskij. Subdirectly irreducible state-morphism BL-algebras. Arch. Math. Logic,
50(1-2):145–160, 2011.
[10] R. Engelking. Dimension theory. North-Holland, 1978.
[11] F. Esteva, L. Godo, P. Hájek, and F. Montagna. Hoops and fuzzy logic. J. Logic Comput.,
13(4):531–555, 2003.
[12] G. Ewald. Combinatorial convexity and algebraic geometry. Springer, 1996.
[13] M. Fedel, K. Keimel, F. Montagna, and W. Roth. Imprecise probabilities, bets and functional
analytic methods in Lukasiewicz logic. To appear in Forum Mathematicum, 2009.
[14] K. R. Goodearl. Partially ordered abelian groups with interpolation. American Mathematical
Society, Providence, R.I., 1986.
[15] P. Hájek. Metamathematics of fuzzy logic, volume 4 of Trends in logic. Kluwer, 1998.
24 G. PANTI, D. RAVOTTI

[16] W. Hurewicz and H. Wallman. Dimension theory. Princeton Mathematical Series, v. 4.


Princeton University Press, 1941.
[17] A. I. Kokorin and V. M. Kopytov. Fully ordered groups. Wiley, 1974.
[18] T. Kroupa. Every state on semisimple MV-algebra is integral. Fuzzy Sets and Systems,
157(20):2771–2782, 2006.
[19] L. Kuipers and H. Niederreiter. Uniform distribution of sequences. Dover, 2006. First pub-
lished in 1974 by Wiley-Interscience.
[20] V. Marra. The Lebesgue state of a unital abelian lattice-ordered group. II. J. Group Theory,
12(6):911–922, 2009.
[21] G. Metcalfe, N. Olivetti, and D. Gabbay. Proof theory for fuzzy logics, volume 36 of Applied
Logic Series. Springer, 2009.
[22] D. Mundici. Averaging the truth-value in Lukasiewicz logic. Studia Logica, 55(1):113–127,
1995.
[23] D. Mundici. The Haar theorem for lattice-ordered abelian groups with order-unit. Discrete
Contin. Dyn. Syst., 21(2):537–549, 2008.
[24] D. Mundici. Advanced Lukasiewicz calculus and MV-algebras, volume 35 of Trends in Logic—
Studia Logica Library. Springer, 2011.
[25] M. R. Murty. Problems in analytic number theory, volume 206 of Graduate Texts in Mathe-
matics. Springer, 2001.
[26] G. Panti. The automorphism group of falsum-free product logic. In S. Aguzzoli et al., editor,
Algebraic and Proof-theoretic Aspects of Non-classical Logics, number 4460 in Lecture Notes
in Artificial Intelligence, pages 275–289. Springer, 2007.
[27] G. Panti. Invariant measures in free MV-algebras. Comm. Algebra, 36(8):2849–2861, 2008.
[28] G. Panti. Denominator-preserving maps. Aequationes Math., 84(1-2):13–25, 2012.
[29] C. P. Rourke and B. J. Sanderson. Introduction to piecewise-linear topology. Springer, 1972.
[30] W. Rudin. Real and complex analysis. McGraw-Hill, third edition, 1987.
[31] R. P. Stanley. Enumerative combinatorics. Vol. 1, volume 49 of Cambridge Studies in Ad-
vanced Mathematics. Cambridge University Press, 1997. With a foreword by Gian-Carlo Rota.
Corrected reprint of the 1986 original.
[32] P. Stein. A note on the volume of a simplex. Amer. Math. Monthly, 73(3):299–301, 1966.
[33] K. Yosida. On the representation of the vector lattice. Proc. Imp. Acad. Tokyo, 18:339–342,
1942.
[34] G. M. Ziegler. Lectures on polytopes, volume 152 of Graduate Texts in Mathematics. Springer,
1995.

Department of Mathematics, University of Udine, via delle Scienze 206, 33100 Udine,
Italy

You might also like